You are on page 1of 26

ARTICLE IN PRESS

Prog. Polym. Sci. 31 (2006) 9831008


www.elsevier.com/locate/ppolysci

Addition polymers from natural oilsA review


Vinay Sharma, P.P. Kundu
Department of Chemical Technology, Sant Longowal Institute of Engineering and Technology, Sangrur, Punjab 148106, India
Received 23 January 2006; received in revised form 14 September 2006; accepted 15 September 2006

Abstract
Emerging technological knowledge is leading research into new ventures. One such is the conversion of natural oils to
polymers to augment the use of petroleum products as the source of polymeric raw materials. Natural oils, such as vegetable
oils, now mainly used in the food industry, offer alternatives, and recent research has studied new routes of synthesis of
polymers from natural oils. This review paper discusses the synthesis and characterization of new polymers from different
natural oils such as soybean, corn, tung, linseed, castor, and sh oil. The effects of different levels of unsaturation in the
natural oils and various types of catalysts and comonomers on the properties of copolymers are considered.
r 2006 Elsevier Ltd. All rights reserved.
Keywords: Natural oils; Dynamic mechanical analysis; Cross-linking; Polymerization; Drying oil; Glass transition temperature

Contents
1.
2.

3.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 984
Polymers from natural oils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 985
2.1. Soybean oil polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 985
2.1.1. Unmodied soybean oil polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 985
2.1.2. Modied soybean oil polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 992
2.2. Fish oil polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 994
2.3. Corn oil polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 995
2.4. Tung oil polymers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 996
2.5. Linseed oil polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 998
2.5.1. Natural linseed oil polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 998
2.5.2. Epoxidized linseed oil polymers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1002
2.6. Castor oil polymers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1003
2.7. Polymers from other oils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1004
Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1005
Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1006
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1006

Corresponding author. Tel.: +91 16 7283606; fax: +91 16 7283657.

E-mail address: ppk923@yahoo.com (P.P. Kundu).


0079-6700/$ - see front matter r 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.progpolymsci.2006.09.003

ARTICLE IN PRESS
V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

984

1. Introduction
In recent years natural oils have attracted
renewed attention as raw materials for the preparation of resins and polymeric materials, to replace or
augment the traditional petro-chemical based polymers and resins. Natural oils such as linseed and
tung oil have long found various uses in the paint
and varnishes industries. These oils have traditionally been used in organic coatings either as resins or
as a raw material for the preparation of resins.
Soybean oil, safower oil, sunower oil and canola
oil have also been used in polymerizations.
Natural oils are tri-glyceride esters of fatty acids,
the general structure of which is shown in Fig. 1.
Triglycerides comprise three fatty acids joined by a
glycerol center [1]. Most of the common oil contains
fatty acids that vary from 14 to 22 carbons in
length, with 13 double bonds. The fatty acid
distribution of several common oils is shown in
Table 1 [1]. In addition, there are some oils comprise
fatty acids with other types of functionalities (e.g.,

epoxies, hydroxyls, cyclic groups and furanoid


groups) [2]. It is apparent that on a molecular level,
these oils are composed of many different types of
triglyceride, with numerous levels of unsaturation.
In addition to their application in the food industry,
triglyceride oils have been used for the production
of coatings, inks, plasticizers, lubricants and agrochemicals [39]. In general, drying oils (these can
polymerize in air to form a tough elastic lm) are
the most widely used oils in these industries,
although the semi-drying oils (these partially harden
when exposed to air) also nd use in some
applications. The polymers obtained from natural
oils are biopolymers in the sense that they are
generated from renewable natural sources; they are
often biodegradable as well as non-toxic.
Some biopolymers obtained from natural oils are
exible and rubbery. Generally, they are prepared
as cross-linked copolymers. Bacterial polyesters are
obtained from a large number of bacteria when
subjected to metabolic stress. The cross-linking
process for unsaturated bacterial polyester is shown

fatty acid chain


glycerol center

oleic acid chain

O
O

linoleic acid chain

linolenic acid chain

O
O

three ester bonds

Fig. 1. The triglyceride chain containing three fatty acid chains joined by a glycerol center. Reprinted with permission from Polymer 2001;
42: 1569 r Elsevier Science Ltd., [10].

Table 1
Main fatty acid contents in different oils
Fatty acid

[#C: #DB] Canola oil Corn oil Cottonseed oil Linseed oil Olive oil Soybean oil Tung oil Fish oily

Palmitic
Stearic
Oleic
Linoleic
Linolenic
a-elaeostearic acid
Average #DB/triglyceride.

16:0
18:0
18:1
18:2
18:3

4.1
1.8
60.9
21.0
8.8

3.9

10.9
2.0
25.4
59.6
1.2

4.5

21.6
2.6
18.6
54.4
0.7

3.9

5.5
3.5
19.1
15.3
56.6

6.6

13.7
2.5
71.1
10.0
0.6

2.8

11.0
4.0
23.4
53.3
7.8

4.6

4
8
4

84
7.5

18.20
1.10
0.99

3.6

Reproduced with the permission from J Appl Polym Sci 2001; 82: 703 r John Wiley and Sons, Inc. [1].
 #C stands for number of carbon atoms in chain and #DB stands for the number of double bonds in that chain.
y
Fish oils tend to contain a high double bond content; for example, the composition of a Norway sh oil examined in one study
contained a fatty acid (ethyl ester) composition with 8.90% having no double bonds, 6.03% having four double bonds, 37.25% having
EPA or DPA and 24.72% (DHA) having six double bonds [29].

ARTICLE IN PRESS
V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

with cyclohexane dicarboxylic acid. Scheme 2(b)


shows the oligomerization with maleic acid, which
introduces more double bonds in the oligomers.

in Fig. 2, describing the cross-linking of the


unsaturated bacterial polyesters prepared from
soybean oily fatty acids. It is observed that crosslinking occurs in at least two polyester chain double
bonds. Scheme 1 describes the representative
process of cationic copolymerization of the triglyceride oil with styrene and divinylbenzene in the
presence of a modied boron triuoride diethyl
etherate. Scheme 2 shows the oligomerization of a
modied acrylated epoxidized soybean oil (AESO)
with reagents selected to stiffen the polymer chain.
Scheme 2(a) shows the oligomerization of an AESO

2. Polymers from natural oils


2.1. Soybean oil polymers
2.1.1. Unmodified soybean oil polymers
Polymers derived from soybean oils have been
extensively investigated by Larock et al. [1015];
soybean oils are biodegradable vegetable oil, readily
O

O
CH

CH2

CH

CH2

CH2

HC

CH2

O
CH

CH

985

CH2

CH

CH2

Fig. 2. The cross-linking process of bacterial polyester obtained from soybean oily fatty acids. Reprinted with the permission from Polym
Bull 2001; 46: 393 r Springer-Verlag, Inc. [24].

CH
CH
CH

CH

CO2

CH

CO2

CH

CO2

O
H2C

O
C
CH2

CH

HC
m

CH

CH

O
O

C
CH

CH

O
CH2

CH
n

CH

CH2

CH

Scheme 1. The proposed process of cross-linking of natural oil with styrene and divinylbenzene in presence of modied initiator.
Reprinted with the permission from J Appl Polym Sci 2003; 90: 1832 r Wiley Periodicals, Inc. [30].

ARTICLE IN PRESS
986

V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

Scheme 2. The modication of acrylated epoxidized soybean oil (AESO) shown using cyclohexane dicarboxylic anhydride or maleic
anhydride. These AESOs were cured with styrene or other comonomers. Reprinted with the permission from J Appl Polym Sci 2001; 82:
707 r John Wiley and Sons, Inc. [1].

available in bulk; specication of soybean oils used


by the Larock group are reported in Table 2.
Natural soybean oil possesses a triglyceride structure with highly unsaturated fatty acid side chains.
The 1H NMR spectra of some example oils are
shown in Fig. 3. The unsaturation in these oils
makes them ideal monomers for the preparation of

various polymers. Cationic copolymerization of


regular soybean oil, low saturated soybean oil or
conjugated low saturated soybean oil with styrene
and divinylbenzene leads to various copolymers.
These copolymers have been characterized by
various techniques, including dynamic mechanical
analysis (DMA), thermogravimetric analysis

ARTICLE IN PRESS
V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

987

Table 2
The composition of the soybean oils used for the preparation of copolymers
Soybean oil

Regular soybean oil


Low saturated soybean oil
Conjugated saturated soybean oil

Fatty acidsb

CQQC
Type

No.a

C16:0

C18:0

C18:1

C18:2

C18:3

Non-conjugated
Non-conjugated
Conjugated

4.5
5.1
5.1

10.5
5.0
5.0

3.2
3.0
3.0

22.3
20.0
20.0

54.4
64.0
64.0

8.3
9.0
9.0

Reproduced with the permission from J Polym Sci: Part B: Polym Phys 2000; 38: 2722 r John Wiley and Sons, Inc. [12].
a
The average number of carbon-carbon double bonds was calculated by 1H NMR spectral analysis.
b
For example, C18:2 represent the fatty acid (ester) that possesses 18 carbons and 2 CQQC bonds.

Fig. 3. The 1H NMR spectra of different soybean oils. (a) regular soybean oil, (b) low saturated soybean oil and (c) conjugated low
saturated soybean oil. Reprinted with permission from J Appl Polym Sci 2001; 80: 660 r John Wiley and Sons, Inc. [11].

(TGA), differential scanning calorimetry (DSC),


scanning electron microscopy (SEM) and thermal
mechanical analysis (TMA).
Cationic polymerization of the soybean oil with
divinylbenzene comonomer initiated by boron triuoride diethyl etherate results in polymers ranging
from soft rubbers to hard thermosets, depending on
the oil and the stoichiometry employed [10]. It was
found that the initiator was immiscible with these
oils, but that miscibility was vastly improved when
the initiator was modied with a norway sh oil
ethyl ester. The copolymerization of soybean oil
with styrene and norbornadiene or dicyclopentadiene initiated by boron triuoride diethyl etherate
resulted in polymers with good mechanical properties and thermal stability.
It has been observed that the copolymerization of
soybean oils with other comonomers results in a
network, with a gelation time dependent on the

stoichiometry and type of the triglyceride oil used


[11]. The gelation time and yield for various
copolymers prepared from varying concentrations
of the oils, comonomers and modied initiators is
reported in Table 3. The yield of the cross-linked
product depends on the concentration of the crosslinking agents, such as divinylbenzene, dicyclopentadiene, etc. As usual, cross-linking increases the
glass transition temperature of the polymer. Polymers from different soybean oils show different
properties, and the cross-linking density of the bulk
polymers considerably affected their thermophysical
properties [12]. Several copolymers obtained from
copolymerization of a soybean oil with divinylbenzene were characterized by DMA, TGA and soxhlet
extraction, with the results shown in Tables 4 and 5.
From these results, it was clear that the composition
of the copolymer dictated the properties. For
example, the oily component of the copolymer

ARTICLE IN PRESS
V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

988

Table 3
The results from copolymerization of different soybean oils using different modied initiator system with styrene and divinylbenzene,
norbornadiene or dicyclopentadiene
Original composition (wt%)
Triglyceride oil

Comonomers

Initiators

45%
45%
45%
45%
45%
45%
45%
45%
45%
45%

32%ST+15%DVB
32%ST+15%DVB
32%ST+15%DVB
32%ST+15%DVB
32%ST+15%DVB
32%ST+15%DVB
32%ST+15%DVB
32%ST+15%DVB
32%ST+15%NBD
32%ST+15%DCP

5%SG-I+3%BFE
5%SG-II+3%BFE
5%SG-III+3%BFE
5%NFO+3%BFE
5%NFO+3%BFE
5%NFO+3%BFE
5%NFO+3%BFE
5%NFO+3%BFE
5%NFO+3%BFE
5%NFO+3%BFE

LSS
LSS
LSS
LSS
SOY
LSS
CLS
CLS
CLS
CLS

Gelation time (s)

Yield (%) of cross linked


polymer after extraction

3.0  102
3.0  102
3.0  102
3.0  102
2.4  102
3.0  102
6.6  102
6.6  102
3.5  103
2.1  105

83
82
83
84
80
84
92
92
89
80

Reproduced with the permission from J Appl Polym Sci 2001; 80: 662 r John Wiley and Sons, Inc. [11].

Table 4
The DMA, TGA, and Soxhlet extraction results for the samples prepared by copolymerization of soybean oil and divinylbenzene in the
presence of modied initiator
Polymer samplea

SOY60-DVB35-BFE5
SOY50-DVB35-(NFO10-BFE5)
SOY55-DVB30-(NFO10-BFE5)
SOY60-DVB25-(NFO10-BFE5)
LSS60-DVB35-BFE5
LSS50-DVB35-(NFO10-BFE5)
LSS55-DVB30-(NFO10-BFE5)
LSS60-DVB25-(NFO10-BFE5)
CLS50-DVB35-(NFO10-BFE5)
CLS55-DVB30-(NFO10-BFE5)
CLS60-DVB25-(NFO10-BFE5)

Eroom(Pa)  108

4.0
5.0
2.5
1.7
6.0
7.0
3.8
1.9
12
10
7.8

ne(mol/m3)  103

7.60
11.6
6.51
4.18
10.4
13.0
8.35
4.18
18.9
11.4
7.21

Tg(1C)

Structure (wt%)

TGA (1C)
b

a1

a2

Cross-linked

Free oil

Inc. oil

T10

T50

27
70
15
20
37
70
30
17
90
80
68

10
5
0
0
8
0
-

69
77
75
73
82
84
80
77
88
86
86

31
23
25
27
18
16
20
23
22
14
14

29
37
40
43
42
44
45
47
48
51
56

415
425
380
360
423
425
405
395
440
436
433

490
491
475
470
485
486
486
485
485
486
483

Reproduced with the permission from Polymer 2001; 42: 1573 r Elsevier Science Ltd., [10].
Eroom Youngs modulus at room temperature.
ne Cross-linking density.
a
Here SOY represents regular soybean oil, LSSLow saturated soybean oil, CLSconjugated low saturated soybean Oil, DVB
divinylbenzene, NFONorway Pronova sh oil ethyl ester and BFE boron triuoride diethyl etherate. The numerals, such as SOY60
represents 60 wt% of soybean oil.
b
Wt% of oil incorporated into the cross-linked network.

induces reduction in the glass transition temperature, stiffness and modulus.


The variation of the storage modulus (E0 ) and loss
factor (tan d) with temperature is shown in Figs. 4
and 5, respectively, for several copolymers prepared
from regular soybean oil. In Fig. 4, E0 is minimum
for regular soybean oil and maximum for conjugated low saturated soybean oil. In Fig. 5, the
polymers from these oils exhibited a single loss peak

at temperatures dependent on the oil, e.g., 68 1C for


regular soybean oil, 61 1C for low saturated soybean
oil and 76 1C for conjugated low saturated soybean
oil. This single loss peak indicates that the polymers
had a homogeneous phase. From these results, it is
clear that in all soybean oils used, the conjugated
low saturated soybean oil gave the highest crosslinking density, glass transition and storage moduli.
Copolymers were also prepared from soybean oil

ARTICLE IN PRESS
V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

989

Table 5
The tensile test results for various soybean oils
Polymer sample

Tg (1C)

ne (mol/m3)  102

E (mpa)

sb (mpa)

eb (%)

Toughness (mpa)

SOY45st32-DVB15-(NFO5-BFE3)
LSS45st32-DVB15-(NFO5-BFE3)
CLS45st32-DVB15-(NFO5-BFE3)

68
61
76

1.8
5.3
22

71
90
225

4.1
6.0
11.5

57.1
64.1
40.5

1.67
2.86
4.00

Reproduced with the permission from J Polym Sci: Part B: Polym Phys 2000; 39: 62 r John Wiley and Sons, Inc. [13].
Tg Glass transition temperature.
ne Cross-linking density.
E Youngs modulus.
sb Ultimate tensile strength.
eb Elongation at break.

1.5

1. E+10

SOY45-ST32-DVB15-(NFO5-BFE3)
LLS45-ST32-DVB15-(NFO5-BFE3)
CLS45-ST32-DVB15-(NFO5-BFE3)

1. E+08
Tan

Storage Modulus (Pa)

1. E+09

1. E+07

0.5

1. E+06
SOY45-ST32-DVB15-(NFO5-BFE3)
LSS45-ST32-DVB15-(NFO5-BFE3)
CLS45-ST32-DVB15-(NFO5-BFE3)

1. E+05
-35

-15

25

45

65

85

105

125

Temperature (C)

Fig. 4. The temperature dependence of the storage modulus (E0 )


on the copolymers prepared from regular soybean oil (SOY),
Lowsat soy oil (LSS) and conjugated Lowsat soy oil (CLS) with
styrene (ST) and divinylbenzene (DVB), using Norway sh oil
modied initiator. Reprinted with permission from J Polym
Sci: Part B: Polym Phys 2000; 38: 2726 r John Wiley and Sons,
Inc. [12].

and divinylbenzene, using boron triuoride diethyl


etherate, resulting in heterogeneous polymeric
materials [10].
The tensile properties of several soybean oil
polymers ranging from elastomers to hard, ductile
and relatively brittle polymers are shown in Fig. 6
[13]. Generally, it is observed that the ultimate
tensile strength increases and the elongation at
break decreases with an increase in the degree of
cross-linking. At lower strain (o10%), the increase
in stress is rapid, while at higher strain (410%), the
regular and low saturated soybean oil polymers
exhibit a slow increase in the stress. The conjugated

0
-35

-15

25
45
65
Temperature

85

105

125

Fig. 5. The temperature dependence of the loss modulus (tan d)


for the copolymers prepared from regular soybean oil (SOY),
Lowsat soy oil (LSS) and conjugated Lowsat soy oil (CLS) with
styrene (ST) and divinylbenzene (DVB), using Norway sh oil
modied initiator. Reprinted with permission from J Polym
Sci: Part B: Polym Phys 2000; 38: 2727 r John Wiley and Sons,
Inc. [12].

low saturated soybean oil polymers show a yield


point.
The tensile fracture surface of polymer samples
(with 35 weight % conjugated low saturated
soybean oil) was observed under a scanning electron
microscope and shown to be very similar to those of
epoxies [Fig. 7]. The SEM micrograph of the
fracture surface and the mist region of the fractured
surface are shown in Figs. 7(a) and 7(b), respectively, for one sample.
The results for damping properties of several
soybean oils over a broad range of temperature and
frequency are reported in Table 6 [14]. The high

ARTICLE IN PRESS
V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

990

SOY45-ST32-DVB15-(NFO5-BFE3)
LLS45-ST32-DVB15-(NFO5-BFE3)
CLS45-ST32-DVB15-(NFO5-BFE3)

Stress (MPa)

12

0
0

20

40
Strain (%)

60

80

Fig. 6. The tensile stress-strain curves from three soybean oil


polymers i.e. regular (SOY), low saturated (LSS) and conjugated
low saturated (CLS) soybean oil polymers for same percentage of
oil and comonomers styrene (ST) and divinylbenzene (DVB).
Reprinted with permission from J Polym Sci: Part B: Polym Phys
2000; 39: 63 r John Wiley and Sons, Inc. [13].

damping intensities are ascribed to the contribution


from the large number of ester groups directly
attached to the soybean oilstyrenedivinylbenzene
copolymer chains. The variation in the glass
transition temperature with cross-linking density is
shown in Fig. 8 for several soybean oil polymers.
The three soybean oil polymers showed the same
glass transition temperature, but differ in the value
of the loss tangent maxima. The broad damping
regions were attributed to segmental inhomogeneity
induced on cross-linking. However, cross-linking
also reduced the damping intensities by restricting
the polymer segmental motions of the homogeneous
polymeric materials. Thus, it is expected that
efcient damping materials (for sound and vibrational applications) would result on the chemical
or physical combination of two or more structurally dissimilar soybean oil-based polymers
to form interpenetrating networks (IPN) with a
phase separated morphology. In such a case,
broad damping regions would be facilitated by
phase microheterogeneity resulting from the formation of IPNs, rather than by the segmental
inhomogeneity [14].
Some soybean oil polymers prepared by cationic
copolymerization show a good shape-memory effect
[15]. Shape-memory refers to the ability of some

Fig. 7. The SEM micrograph of sample CLS35ST39-DVB18(NFO5-BFE3) highlighting the mechanically fractured surface of
the sample and the mist region of the mechanically fractured
surface. Reprinted with permission from J Polym Sci: Part B:
Polym Phys 2000; 39: 75 r John Wiley and Sons, Inc. [13].

materials to remember a specic shape on demand,


even after very severe deformation. Such materials
have applications in civil construction, mechanics
and manufacturing, electronics and communications, printing and packaging, medical equipment,
recreation and sports, and household items.
A shape-memory polymer exhibits mechanical
behavior that includes xing the deformation of
the plastics at room temperature and recovering the
deformation as elastomers at relatively high temperatures [1619]. Shape-memory polymers basically consist of two phases: a reversible phase and a
xed phase. The reversible phase refers to the
polymer matrix, which has a glass transition
temperature (Tg) or a melting temperature (Tm)
well above the application temperature. The xed
phase is composed of either chemical or physical
cross-links that are relatively stable at a temperature
higher than the Tg or Tm of the reversible phase.
At a temperature above Tg or Tm, the shapememory polymer achieves a rubbery elastic state in

ARTICLE IN PRESS
V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

991

Table 6
Results for the damping properties of the copolymers prepared from different soybean oils
Polymer sample

Tg(1C)

ne(mol/m3)

(tan d)max

(tan d)rt

DT at tan d40.3 (1C)

TA (K)

Half-width (1C)

SOY35st39-DVB18-(NFO5-BFE3)
SOY45st32-DVB15-(NFO5-BFE3)
SOY55st25-DVB12-(NFO5-BFE3)
LSS35st39-DVB18-(NFO5-BFE3)
LSS45st32-DVB15-(NFO5-BFE3)
LSS55st25-DVB12-(NFO5-BFE3)
CLS35st39-DVB18-(NFO5-BFE3)
CLS45st32-DVB15-(NFO5-BFE3)
CLS55st25-DVB12-(NFO5-BFE3)

79
68
30
80
61
32
82
76
38

4.7  102
1.8  102
1.0  102
7.3  102
5.3  102
3.9  102
3.4  103
2.2  103
6.5  102

0.88
0.85
0.84
0.86
0.89
1.00
0.94
0.79
1.08

0.12
0.32
0.83
0.11
0.37
0.96
0.07
0.18
0.80

52115 (63)
23113 (90)
265 (67)
23113 (90)
1997 (78)
683 (89)
58116 (58)
48120 (72)
1077 (67)

37.5
48.4
36.3
48.4
46.2
50.1
41.8
43.1
52.9

47
61
51
51
52
57
42
53
44

Reproduced with permission from Polymers for Advanced Technologies 2002; 13: 439,441 r John Wiley and Sons, Ltd., [14].
Tg Glass transition temperature.
ne Cross-linking density.
(tan d)max Loss tangent maxima.
(tan d)rt Loss tangent at room temperature.
TA tan d area.

120

100

Tg (C)

80

60

40

20

SOY45-(ST+DVB) 47-(NFO5-BFE3)
LSS45-(ST+DVB) 47-(NFO5-BFE3)
CLS45-(ST+DVB) 47-(NFO5-BFE3)

0
10

100

1000
e

10000

100000

(mol/ m3)

Fig. 8. The dependence of the glass transition temperature (Tg)


on cross-linking density (ne) for different soybean oil polymers.
Reprinted with permission from Polym Adv Technol 2002; 13:
444 r John Wiley and Sons, Ltd. [14].

which it can be easily deformed by an external force.


When the polymer is cooled to room temperature,
the deformation is xed due to the frozen micro
Brownian motion of the reversible phase. The
hardened reversible phase effectively resists elastic
recovery resulting from the tendency of the ordered
chains to return to a more random state, but the
deformed shape readily returns to its original shape
upon heating above Tg or Tm. The driving force for
the shape recovery is primarily entropy, especially

the strong relaxations of the oriented polymer


chains between the cross-links.
Table 7 shows the shape-memory properties of
several soybean oil polymers. It is observed that the
type of soybean oil greatly affects the shapememory properties of these polymers (e.g., see
No. 13). The polymers from reactive soybean oil
show higher degree of xed deformation (FD value)
and a lower deformability at a temperature higher
than Tg (D value). All of the polymeric materials
show 100% recovery (R) of xed deformation upon
reheating to Tg plus 50 1C.
The time for gelation and vitrication of various
soybean oil polymer systems has been investigated
over a range of isothermal curing temperatures [20].
All the fully cured thermosets were rst made at
room temperature and then subjected to post-curing
at elevated temperatures. The thermal stability and
dynamic mechanical behavior of the resulting
thermosets were not particularly sensitive to the
curing conditions. However, varying the curing time
at low and high temperatures did affect the
structural characteristics of the polymer backbone,
affecting the shape-memory and tensile mechanical
properties. The isothermal timetemperaturetransformation (TTT) cure diagram, developed to study
the epoxy systems [2123], is a very useful tool for
investigating the cure process of the soybean oil
systems. The cure temperatures between Tg,gel, (Tg
at which the system gels and vitries simultaneously) and TgN (maximum Tg of fully cured
system), where gelation precedes vitrication, are of
practical importance. It was observed that gelation

ARTICLE IN PRESS
992

V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

Table 7
Shape memory properties of soybean oil polymers
Polymer sample

SOY45ST32-DVB15-(NFO5-BFE3)
LSS45ST32-DVB15-(NFO5-BFE3)
CLS45ST32-DVB15-(NFO5-BFE3)
SOY45ST32-(DVB5-NBD5-DCP5)-(NFO5-BFE3)
(SOY22.5-LSS22.5)-ST32-(DVB5-NBD5-DCP5)-(NFO5-BFE3)
(SOY15-LSS15-CLS15)-ST32-(DVB5-NBD5-DCP5)-(NFO5-BFE3)
SOY45ST20-(DVB9-NBD9-DCP9)-(NFO5-BFE3)
(SOY22.5-LSS22.5)-ST20-(DVB9-NBD9-DCP9)-(NFO5-BFE3)
(SOY15-LSS15-CLS15)-ST20-(DVB9-NBD9-DCP9)-(NFO5-BFE3)

Tg (1C)

68
61
76
42
43
44
68
70
74

ne (mol/m3)

1.8  102
5.3  102
2.2  103
9.8  10
1.3  102
2.7  102
3.1  102
3.7  102
5.2  102

Shape memory results (%)


D

FD

100
86
77
100
100
100
100
100
100

80
96
98
63
74
75
97
98
99

100
100
100
100
100
100
100
100
100

Reproduced with permission from Journal of Applied Polymer Science 2002; 84: 1539 r John Wiley and Sons, Ltd., [15].
Tg Glass transition temperature.
ne Cross-linking density.
D Deformability of the material at temperature higher than Tg.
FD Degree to which the deformation is xed at ambient temperature.
R Final shape recovery.

Fig. 9. The 1H NMR spectra of acrylated epoxidized soybean oil (AESO). Reprinted with the permission from J Appl Polym Sci 2001;
82: 710 r John Wiley and Sons, Inc. [1].

occurs at approximately 15% conversion of the


reactants, and the yield of cross-linked polymers
continued to increase following gelation. However,
only about 50% of the soybean oil reactants were
converted into cross-linked polymers when the
system vitried. Thus, in order to obtain fully cured
networks, the materials were subsequently postcured at elevated temperatures.
2.1.2. Modified soybean oil polymers
Several types of functionalization can be obtained
at various active sites within the triglyceride
structure, such as the double bond, the ester group,
the allylic carbons, and the carbons a to the ester

group. Various chemical pathways for functionalization of these triglycerides have been studied.
LaScala and Wool et al. [1] analyzed optimization
of the effect of chemical functionalization on the
mechanical properties and thermal stability of some
resins. The viscoelastic properties of resin samples
made of AESO and cured at room temperature with
varying amounts of styrene were studied. The 1H
NMR spectra of AESO are shown in Fig. 9. Some
triglyceride-based monomers prepared from acrylic
acid are shown in Fig. 10. It was found that both E0
and Tg increased with increasing styrene content in
the copolymers. The styrene content at 33.3 wt%, or
a 2:1 AESO to styrene ratio, was considered optimal

ARTICLE IN PRESS
V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

993

O
O

OH

O
O

OH

OH

OH
O

O
O
O

Acrylated Epoxidized Soybean Oil (AESO)

OH

O
O

O
O
O
O
O

OH

Maleinated Soybean Oil Monoglyceride (SOMG/MA)


O

HO
O

O
OH
O

HO
O

OH

O
O

OH

OH

HO
O

O
O

HO
O

Maleinated Hydroxylated Soybean Oil (HSO/MA)

Fig. 10. Different triglyceride-based acrylic monomers. Reprinted with the permission from J Appl Polym Sci 2001; 82: 706 r John Wiley
and Sons, Inc. [1].

with respect to the properties. The variations of


damping peak (tan d) of AESOstyrene copolymer
with temperature are shown in Fig. 11. The different
compositions of the AESOstyrene copolymers
were used to study the dynamic mechanical
behavior of the copolymers. The copolymer with
50% AESO showed a very sharp loss tangent peak,
which broadened with the increasing contents of
AESO.
Baki Hazer and co-workers [24] reported that the
soybean fatty acid and poly(hydroxy alkanoate)
(PHA) reacted under UV irradiation to form a
cross-linked biopolyester. It is believed that in the

prevailing conditions, the esterication reaction


proceeds along with cross-linking via a free radical
mechanism. A bacterial polyester containing olenic
groups in the side chains was prepared by feeding
pseudomonas oleovorans with soybean fatty acids.
The structure of PHA with unsaturated side chains
soybean fatty acids is shown in Fig. 12. After crosslinking, the biopolyester became a smooth and less
sticky elastomeric lm. The cross-linked biopolyester, obtained thermally at 60 1C with benzoyl
peroxide initiation, exhibited the highest crosslinking density. The cross-linking was determined
by solgel analysis. The swelling properties of these

ARTICLE IN PRESS
V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

994

side chain [26]. Fish oils nd use in industry in the


production of protective coatings, lubricants, sealants, inks, animal feed and surfactants [27].
Regular and conjugated sh oils prove to be
cationically polymerizable monomers [26]. Polymers
from sh oil have been prepared with divinylbenzene, dicyclopentadiene and norbornadiene comonomers [26]. The polymers obtained from sh oil
were found to be typical thermosetting materials
[26]. Three distinct decomposition states were found
by TGA, corresponding to the evaporation of
unsaturated free oil present in the bulk polymer,
carbonization of the cross-linked polymer and
subsequent oxidation of the carbons [26].
Both regular oil and conjugated sh oils can be
used to produce hard and stiff polymers having
moduli comparable to those of the conventional
plastics [28]. In Fig. 13, sh oil polymers are
compared with commercially available polymers.
The glass transition temperature of the conjugated
sh oil (CFO) polymer is nearly 110 1C, which is
higher than Tg for polystyrene. The thermosetting
nature of the sh oil polymers results in an
improvement in thermal stability at higher temperatures (T4200 1C) as compared to the thermoplastic
polymers. The conjugated sh oil resulted in
polymeric materials with better thermal stability
and mechanical properties as compared to the
regular sh oil polymers. However, regular sh oil
polymers were observed to have relatively good
creep resistance properties [28]. These polymers
were found to have high cross-linking densities. The
bulk polymer was mainly composed of insoluble
substances (6580 wt%). The homopolymerization
of regular sh oil ethyl ester initiated by boron
triuoride diethyl etherate (BFE) resulted in a

biopolyesters were studied and the number average


molecular weights were also calculated using the
Flory-Rehner equation [25]. The cross-link density
can be increased by irradiation with shorter
wavelength. It was also reported that while polymerization of the soybean oil is difcult, the
biopolyester from soybean oil fatty acids is more
reactive than the relative oil for polymerization.
2.2. Fish oil polymers
Fish oil is biodegradable and is readily available
as a byproduct in the production of shmeal. Fish
oils have a triglyceride structure with a high
percentage of polyunsaturated omega 3 fatty acid
side chains, which can contain as many as 56 nonconjugated carboncarbon double bonds per ester
0.9
0.8
0.7

100% AESO
80% AESO
60% AESO
50% AESO

Tan

0.6
0.5
0.4
0.3
0.2
0.1
-150

-100

-50

50

100

150

200

Temperature (C)

Fig. 11. Plot of dynamic mechanical properties of AESOstyrene


copolymers against temperature using different compositions
(AESO 50100%) of the copolymers. Reprinted with the
permission from J Appl Polym Sci 2001; 82: 714 r John Wiley
and Sons, Inc. [1].

Saturated units:
x = 2 for hexanoate.
x = 4 for octanoate.
x = 6 for decanoate.
Unsaturated units:
y = 0 for -ene.
y = 1 for -diene.
y = 2 for -triene.
z = 1-3.

CH3
CH
HC

CH2
HC
CH3
CH2
O

CH

CH
O

CH2

x
CH2

CH

O
CH2

PHA-Soybean

Fig. 12. The structure of poly(hydroxy alkanoate) (PHA) containing unsaturated side chains of fatty acids, PHAsoybean. Reprinted with
the permission from Polymer Bulletin 2001; 46: 391 r Springer-Verlag, Inc. [24].

ARTICLE IN PRESS
V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008
1.0
1e+10
Polystyrene

0.8

1e+9
0.6
Tan

Storage Modulus (Pa)

Epoxy

CFO-DVB-38

1e+8

Polyethylene

0.4

Polystyrene
Polyethylene

1e+7

0.2

Epoxy

CFO-DVB-38

1e+6
50

100

150

200

0.0
250

Temperature (C)

Fig. 13. The comparison of the storage modulus (E0 ) and loss
factor (tan d) for conjugated sh oil (CFO-DVB-38) and some
commercially available polymers. Reprinted with the permission
from Polymer 2000; 41: 7935 r Elsevier Science Ltd., [25].

viscous oil and they did not gel at all at room


temperature or above. However, conjugated sh oil
is more reactive than regular sh oil and the simple
homopolymerization of conjugated sh oil results in
a soft and solid material. The addition of alkene
comonomers, such as divinylbenzene, norbornadiene, dicyclopentadiene, etc., facilitated gelation of
regular sh oil ester system. The gel time varies from
1 min to a few hours at room temperature depending upon the oil and comonomer contents. Regular
and conjugated sh oil ethyl ester form polymers
with divinylbenzene in the presence of BFE.
Regular sh oil resulted in soft to hard thermosets
depending upon the amount of divinylbenzene,
whereas conjugated sh oil resulted in very hard
thermosets.
Many other alkene comonomers, such as furfural,
benzoquinone, p-mentha-1, 8-diene, furan, vinyl
acetate, etc., were also examined in BFE-initiated
copolymerizations [29]. Copolymerization of these
alkene comonomers with sh oil or conjugated sh
oil generated soft to hard polymers in the presence
of divinylbenzene, norbornadiene or dicyclopentadiene comonomers. A wide range of alkene comonomers resulted in viable solid plastic materials, but
neither copolymerizing sh oil nor conjugated sh
oil could be copolymerized with p-mentha-1, 8diene, methyl methacrylate, isoprene, methyl crotonate, phenol, linalool, furfural, bisphenol A, etc.
Among various Lewis acids, boron triuoride

995

diethyl etherate proved to be the most effective


initiator for copolymerization.
A number of sh oils have also been polymerized
with styrene and divinylbenzene cationically in the
presence of boron triuoride diethyl etherate as
initiator [30]. In addition to the good thermal
stability, and physical and mechanical properties,
the polymeric products also showed even more
promising and valuable properties, such as good
damping and shape-memory properties. The tensile
fractured surfaces of the sh oil polymers have been
studied to obtain SEM micrographs of the fractured
surfaces of sh oil polymers such as those shown in
Fig. 14. It is observed that Norway sh oil (NFO)
polymers have smooth surfaces even at higher
magnications, whereas triglyceride sh oil (TFO)
polymer results in rough surfaces.
2.3. Corn oil polymers
Corn oil is one of the cheapest commercially
available vegetable oils and nds use in food and
livestock feed as well as in the production of
ethanol, which is utilized as a fuel. It has a
triglyceride structure, with approximately 4.1 carboncarbon double bonds per molecule in fatty acid
side chains [31]. The fairly high degree of unsaturation present in corn oil makes it possible to
copolymerize this oil with other monomers. Otherwise, corn and soybean oils have similar chemical
structures, with three fatty acid chains composed of
oleic acid, linoleic acid and linolenic acid [32].
Polymers have been prepared by cationic polymerization of corn oil, styrene and divinylbenzene in
the presence of boron triuoride diethyl etherate
[31]. Corn oil and conjugated corn oil were
effectively copolymerized with comonomers. The
copolymers obtained with conjugated corn oil
showed better mechanical properties and thermal
stability than the corresponding simple corn oil
polymers. The polymers possess a wide variety of
mechanical properties, such as tensile, exural and
compressive strengths, ranging from elastomers to
tough and rigid plastics. The dynamic mechanical
properties of several corn oil polymers are shown in
Fig. 15.
Fig. 16 shows the dependence of the weight
percent of cross-linked polymers on the cure time
and the corn oil contents in the copolymers.
Fig. 16(a) shows that the conjugated corn oil
resulted in polymers with higher cross-linked contents in comparison to the polymers from regular

ARTICLE IN PRESS
V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

996

Fig. 14. The SEM micrograph of the mechanically fractured surfaces of samples NFO49ST33-DVB15-BFE3, CFO49ST33-DVB15-BFE3
and TFO49ST33-DVB15-BFE3. Reprinted with the permission from Polymer 2001; 42: 10144 r Elsevier Science Ltd. [29].

Storage Modulus (Pa)

1.0E+10
COR35-ST39-DVB18-(NFO5-BFE3)
COR45-ST32-DVB15-(NFO5-BFE3)
COR55-ST25-DVB12-(NFO5-BFE3)

1.0E+9
1.0E+8
1.0E+7
1.0E+6

Loss Factor (tan )

1.0E+5

0.6

0.4

0.2

0
-50

50

100

150

200

Temperature (C)

Fig. 15. The temperature dependence of the storage modulus (E0 )


and loss factor (tan d) for corn oil polymers prepared from
different concentrations of corn oil, keeping the styrene and
divinylbenzene contents in same ratio (3:2). Reprinted with the
permission from J Appl Polym Sci 2003; 90: 1833 r Wiley
Periodicals, Inc. [30].

corn oil. Fig. 16(b) indicates that the system


underwent a slow increase in the yield of crosslinked polymer after gelation. It is also observed
that with the same curing time, only a 5 1C increase
in the cure temperature made an appreciable
difference in the yields of cross-linked polymers
(shown in Fig. 16(b)). A plot of the shape recovery
properties of the conjugated and regular corn oil
copolymers at different temperatures is reported in
Fig. 17. It is observed that 50% shape recovery of
deformations was reached at 42 and 44 1C for
conjugated and regular corn oil polymers, respectively, and attained complete recovery for conjugated corn oil polymer at 75 1C and 96% recovery
for regular corn oil polymer at the same temperature. Regular corn oil copolymer is found to possess
superior damping characteristics when compared to
conjugated corn oil copolymer. This is attributed to
the high reactivity of conjugated oil, leading to
densely cross-linked polymers.
2.4. Tung oil polymers
Tung oil is one of the oldest known drying oils
and nds numerous applications in the paint
industry. It is extracted from the seeds of the tung

ARTICLE IN PRESS
Wt % of cross-linked polymer

V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

100

100
35C
40C

80

80

60

60

40

40

20

20

COR-ST-DVB
CCOR-ST-DVB

0
0

(a)

997

100
200
300
Cure Time (min)

400

0
(b)

20
40
60
80
Corn Oil Contents (%)

100

Fig. 16. (a) show the dependence of the weight percent of the cross-linked polymer obtained from COR45ST32-DVB15-(NFO5-BFE3) on
cure time at 35 and 45 1C and; (b) the weight percentage of cross-linked polymers prepared from conjugated and simple corn oil versus
corn oil contents. Reprinted with the permission from J Appl Polym Sci 2003; 90: 1833 r Wiley Periodicals, Inc. [30].

100

Recovery (%)

80
60
Fig. 18. The structure of a-elaeostearic acid. Reprinted with the
permission from Biomacromolecules 2003; 4: 1018 r American
Chemical Society [34].

40
20

COR45-ST32-DVB15-(NFO5-BFE3)
CCOR45-ST32-DVB15-(NFO5-BFE3)

0
20

40

60
Temperature (C)

80

100

Fig. 17. The study of the shape-recovery results for the polymer
from conjugated and regular corn oil for sample with composition OIL45ST32-DVB15-(NFO5-BFE3) at different temperatures. Reprinted with the permission from J Appl Polym Sci 2003;
90: 1836 r Wiley Periodicals, Inc. [30].

tree, with its main constituent a glyceride of


elaeostearic acid with a conjugated triene structure
(Fig. 18). This highly unsaturated, conjugated
system is largely responsible for the rapid polymerization and outstanding drying properties of the oil
[33]. Tung oil has been polymerized by both free
radical and cationic polymerizations. Larock and
coworkers [34,35] studied the cationic and thermal
polymerization of tung oil in detail. Tung oil is
cationically copolymerized with divinylbenzene in
the presence of boron triuoride diethyl etherate,
resulting in hard plastics [34]. The cationic copolymerization of tung oil is found to be very reactive. It
was observed that tung oil is very sensitive to the
cationic initiator boron triuoride diethyl etherate,

and forms an irregular polymeric solid within a few


seconds after the addition of this initiator at room
temperature. The gel time ranged from a few
seconds to 1 min depending, on the oil and other
monomer compositions. The addition of a less
reactive oil, such as soybean oil, increases the gel
time from second to minutes or hours, again
depending upon the oil and other monomer
compositions. The level of conversion of the starting
material into a cross-linked product depends on the
composition of the material. Tung oil polymers
possess very good dynamic mechanical properties as
well as thermal stability at room temperature. These
are thermally stable upto 200 1C and 10% weight
loss was recorded at around 400 1C.
Tung oil can also be copolymerized with styrene
and divinylbenzene by thermal polymerization [35].
The 1H NMR spectra of the tung oil, styrene,
divinylbenzene and extracted soluble contents from
the sample with 50% tung oil is reported in Fig. 19
[35]. These results are used to calculate the oil
content in a particular sample. The thermal polymerization of tung oil was reported to involve
primarily the dimerization of the elaeostearic acid,
producing monocyclic dimeric fatty acid groups at

ARTICLE IN PRESS
998

V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

2.5. Linseed oil polymers

Fig. 19. The 1H NMR spectra of (a) tung oil (TUN), (b) styrene
(ST), (c) divinylbenzene (DVB) and (d) the extracted soluble
substances from the sample TUN50ST20-DVB30 bulk polymer.
Reprinted with the permission from Biomacromolecules 2003; 4:
1021 r American Chemical Society [34].

200 1C or above [3637]. When polymerized at high


temperatures (in a range of 200300 1C), the resulting
products ranged from viscous oils to weak rubbery
lms. Aromatic comonomers have been introduced
to produce viable polymeric materials. A wide
variety of viable transparent polymeric materials
ranging from elastomers to tough and rigid plastics
have been prepared from thermal copolymerization
of the tung oil, styrene and divinylbenzene [35]. The
stoichiometry and the addition of catalysts greatly
affected the thermophysical and mechanical properties of the polymers. The addition of metallic
catalysts proved to be a very effective means to
accelerate the thermal copolymerization leading to
high cross-link densities and improved properties of
the bulk polymers. However, varying the oxygen
uptake and the addition of various peroxides had no
effect on the resulting polymers.
Trumbo and Mote [38] reported the synthesis of
copolymers from tung oil and diacrylate by DielsAlder reaction and studied the properties of the
lms produced from these copolymers using 1,6hexanediol diacrylate and 1,4-butanediol diacrylate.
The copolymers produced were completely soluble.
The molecular weight distributions of the copolymers were broad and multimodal. The lms of the
copolymers were readily prepared and when cured,
exhibited good solvent resistance, high hardness and
good gloss.

2.5.1. Natural linseed oil polymers


Linseed oil, obtained from the linseed seed, is a
fatty acid ester triglyceride and is composed of about
53% linolenic acid, 18% oleic acid, 15% linoleic acid,
6% palmitic acid and 6% stearic acid [39]. It is
traditionally used as a drying oil for surface-coating
applications. To make it a superior drying oil in
terms of lm properties, different olenic monomers,
such as styrene have been copolymerized with linseed
oil [4042]. Linseed oil is polymerized by cationic,
thermal, free radical polymerization, as well as by
oxidative polymerization [43]. The process of autooxidative curing of linseed oil with initiation,
propagation and termination steps is shown in
Scheme 3. It is observed that in the initiation step,
naturally occurring hydroperoxides decompose to
form free radicals. This step can be catalyzed either
by the inclusion of driers (pigments used as catalysts)
or the application of heat. These free radicals react
with antioxidants and after consuming the antioxidants, react with the fatty acid chains of the drying
oil. The propagation then proceeds by the abstraction of the hydrogen atoms present between double
bonds of the methylene groups, which result in the
free radical 1. Radical 1 is resonance stabilized, and
can react with oxygen to form radical 2, as shown in
Scheme 3. The peroxy free radical may be conjugated
or non-conjugated. This can regenerate free radical 1
by abstracting hydrogen from methylene groups.
After this termination, cross-linking proceeds.
The termination results in the formation of structures
3, 4 and 5. Cobalt, lead and zirconium-2-ethylhexanoates are generally used as catalysts for oxidative
polymerization of linseed oil. It is found that
cobalt-2-ethylhexanoate was active during the oxidative step and lead and zirconium catalysts acted
during the polymerization step. Cobaltzirconium
complex gave the best results, and under specied
conditions, zirconium catalyst was more efcient
than lead.
Soucek et al. [44] used DSC to study the autooxidative curing of linseed oil, catalyzed by various
metal catalysts. A manganese drier was used to
catalyze the reaction at the coating surface (top
drier) and a zirconium drier was used to catalyze the
reaction throughout the entire lm thickness
(through drier). It is a common practice to use a
combination of top and through driers. Fig. 20
shows the DSC scans for linseed oil using different
drier systems at 130 1C.

ARTICLE IN PRESS
V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

999

Scheme 3. The autoxidation process for the curing of linseed oil showing initiation, propagation and termination steps of the reaction.
Reprinted with the permission from Prog Org Coat 1998; 33: 220 r Elsevier Science S. A. [42].

Linseed oil has been styrenated for use in


polymerizations to obtain some desired lm properties [45]. There are two methods of styrenation of
oils, dependent on the mode of generation of free
radicals: the classical method and the macromer
method. In the classical method, free radicals are
formed by thermolysis, either in the absence (for
conjugated oils) or in the presence (for nonconjugated oils) of an initiator, such as benzoylperoxide. The degree of conjugation and unsaturation
has a crucial effect on the formation of free radicals
on the oil molecules. The process of styrenation of
linseed oil proceeds through different stages, de-

picted in Scheme 4, showing the process of


styrenation of linseed oil and castor oil. The oils
are rst interesteried and then the macromer is
prepared from the reaction of interesteried product
and acrylic acid. Finally, the macromers are reacted
with styrene. The macromonomers of linseed oil
were prepared by transesterication of methylmethacrylate (MMA) with partial glycerides [46].
Styrenation was achieved via a free radical mechanism using benzoylperoxide as the initiator. Scheme 5
shows the detailed mechanism for the preparation
of partial glycerides and their styrenation. In the
macromer method, the macromer was obtained

ARTICLE IN PRESS
V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

1000

through reaction of hydroxyl-containing oil specimens with vinyl monomer and then this macromer
was homopolymerized or copolymerized with styrene [47]. Semidrying and non-conjugated oils were
mixed with conjugated drying oils to produce
homogeneous styrenated products [48]. When classical and macromer methods were compared, it was
found that macromer technique resulted in homogeneous products in high yield.
The effect of cobalt-2-ethylhexanoate drier on the
oxidative polymerization of linseed oil was investigated [49]. It was found that the cobalt catalysts
accelerated all of the oxidation reactions involved in
the process, and improved the formation of a solid
lm at the surface, but a viscous oil remained under
this surface lm. The addition of cobalt drier only
inuenced the kinetics, but did not alter the reaction
products.

LIN-Mn (0.1) @ 130C


LIN-TIA (0.5) @ 130C
LIN-TIP (0.5) @ 130C

Exotherm

20

40

60
Time (min)

80

100

Fig. 20. Differential scanning calorimetric isothermal exotherms


for linseed oil cured with three different metal catalysts,
manganese drier, titanium (di-isopropoxide) bis (acetyl-acetonate) (TIA) and titanium (IV) isopropoxide (TIP) at 130 1C.
Reprinted with the permission from Progress in Organic Coatings
1996; 28: 254 r Elsevier Science S. A. [43].

OH

OH

O
O

O
O

Linseed Oil

OH

Castor Oil

Interesterification
OH

O
O

OH

OH

OH

O
O

OH

O
O
O

OH

OH

O
O

OH

Other possible isomers

OH

OH

Interesterification Product (IP)

Acrylic Acid (CH2=CH-COOH)


-H2O
O
O

OH

O
O

OH

Styrene
O

O
O

Macromer

OH
O

O
CH

CH CH2

CH
n

Styrenated Oil

Scheme 4. Styrenation of the linseed oil and castor oil through interesterication of oils which react with acrylic acid and nally with
styrene to form styrenated oil. Reprinted with the permission from Macromol Mater Eng 2000; 283: 17 r WileyVCH Verlag GmbH [44].

ARTICLE IN PRESS
V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

1001

O
O

OH

OH

OH

OH
O

OH
O
O

OH

Glycerol

Triglyceride Oil

Mixture of Partial Glyceride

OH
O

OH
O

OH
O
CH3

-CH3OH

H2C

C O
O CH3

O
O

CH3

O
C

CH2

O
O

O
O

O
C

CH3
C CH2

C
O
O O

Benzoyl peroxide

O H3C
C
C

CH2

CH2

CH3

H2C

HC

O
O

CH2

CH3

CH2

CH2 CH

O
O
O

O
O

CH2

H3C
HC

H2C

O
O

O
O

CH3

O
C

Benzoyl peroxide

H3C
C CH2
O
O O
O
O
O
C CH2

CH2 CH

CH2 CH

H3C
Styrenated Oil

Scheme 5. The process of preparation of the partial glyceride and then the styrenation of the partial glyceride through free radical
mechanism in presence of benzoylperoxide. Reprinted with the permission from J Appl Polym Sci 2003; 88: 237475 r Wiley Periodicals,
Inc. [45].

For lms cured with the cobalt drier, the


quantication of the oxidation product shows lower
concentrations of carboxylic acids, ketones and
alcohols [49]. The efciency of peroxide decomposition by the cobalt, zirconium and calcium/zirconium drier was compared by measuring the
variations of peroxide value as a function of
oxidation time. The inuence of different catalysts
and anhydride hardeners on the curing of polymer
networks based on epoxidized linseed oil has been
studied [50]. Generally, tertiary amines and imidazoles were used as catalysts. The variation of loss

factor and storage modulus with temperature for


epoxidized linseed oilcis-1,2,3,6-tetrahydrophthalic anhydride (ELOTHPA) system is depicted in
Fig. 21. A wide range of temperature from 140 to
+220 1C was used for the study. On using imidazole
catalysts, the epoxidized linseed oil was cured to a
maximum extent, leading to high conversion of
anhydrides (nearly double) and increased stiffness.
By using tertiary amine as a catalyst, lower stiffness
was observed in comparison to that with the
imidazole catalyst. When the catalyst content
was increased, the conversion of anhydride was

ARTICLE IN PRESS
V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

1002

observed to be low due to the fast gelling of the


system. On decreasing the catalyst concentration, it
was observed that the anhydride conversion is
maximum leading to good thermal stability and
mechanical properties, such as exural and Youngs
moduli, Tg, etc.
2.5.2. Epoxidized linseed oil polymers
A series of epoxynorbornane-modied linseed oils
were prepared as a function of the norbornane
content and were characterized by various methods
10000

0.4
0.35
0.3
0.25

100

Tan

E (MPa)

1000

0.2
0.15
0.1

10

0.05
1
-140

0
- 80

20
40
Temperature (C)

100

160

Fig. 21. The temperature dependence of loss factor (tan d) and


storage modulus (E0 ) at a frequency of 1 Hz for an epoxidized
linseed oilcis-1,2,3,6-tetrahydrophthalic anhydride (ELO
THPA) system. Reprinted with the permission from Polymer
2000; 41: 8609 r Elsevier Science Ltd., [49].

of analysis, such as FTIR, NMR, etc. [51]. The


cationic photopolymerization of epoxynorbornane
linseed oil and epoxidized linseed oil was studied by
various methods of characterization. The preparation of epoxynorbornane linseed oil and norbornane linseed oil is detailed in Scheme 6. The 1H
NMR spectra of linseed oil, norbornane linseed oil
and epoxynorbornane linseed oil are shown in Fig.
22. Figs. 23 and 24 show the effect of the reactive
and non-reactive diluents on the rate of polymerization [51]. The addition of reactive or non-reactive
diluents reduced the viscosity of the formulations
and was found to have signicant effect on epoxy
conversion and the rate of polymerization of
epoxidized norbornane linseed oil. It was reported
that the photopolymerization of epoxides was
accelerated due to the presence of vinyl ether
[52,53]. This was attributed to the generation of a
large number of propagating cationic species via a
redox reaction between the vinyl ether and the
photoinitiator diaryliodonium salt. The mechanism
for the photoinitiated cationic polymerization
of epoxynorbornane linseed oil is depicted in
Scheme 7.
Soucek and coworkers [54] have studied the
properties of UV-curable hybrid lms derived from
epoxynorbornane linseed oil. Different levels of
epoxynorbornane linseed oil and tetraethylorthosilane (TEOS) were used in the study. The organic

O
O

O
+

7
O

NLO--25

0.76-0. 90 MPa

NLO--50

250 C

NLO--100
H2O2 /Tungstate
60 C
O

O
O

O
7

NLO--25

[a]

7
O

NLO--50

O
O

7
7

O
O

NLO--100

7
ENLO--25

[b]

ENLO--50

ENLO--100

Scheme 6. The reaction process of the preparation of the norbornane linseed oil (NLO) and epoxynorbornane linseed oil (ENLO).
Reprinted with the permission from J Polym Sci: Part A: Polym Chem 2003; 41: 3444 r Wiley Periodicals, Inc. [50].

ARTICLE IN PRESS
V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

1003

Fig. 22. The 1H NMR spectra of (a) linseed oil, (b) norbornane linseed oil and (c) epoxynorbornane linseed oil. Reprinted with the
permission from J Polym Sci: Part A: Polym Chem 2003; 41: 3446 r Wiley Periodicals, Inc. [50].

inorganic hybrid lms are formed either from the


inorganic phase formation within the organic lm
or a simultaneous polymerization of both organic
and inorganic reactive groups. It was also observed
that the incorporation of TEOS oligomers improved
the performance of lms and enhanced tensile
strength, fracture toughness, thermal stability and
general coating properties of epoxynorbornane
linseed oil.
Turri et al. [55] studied the polymerization of
linseed oil via calorimetry. They observed that all of
the pigments such as minium (Pb3O4), chromium
yellow (PbCrO3) and red earth (based on iron
oxides) accelerate the polymerization reaction of
linseed oil and therefore have a catalytic action.
Kundu and Larock prepared a variety of new
polymers from conjugated linseed oil, styrene and
divinylbenzene by thermal polymerization [56]. The
resulting polymeric material was opaque and contained 3585% of cross-linked materials. These
copolymers exhibited a major thermal degradation

of 7290% at 493500 1C. The thermogravimetric


behavior of some samples is shown in Fig. 25. All of
the samples exhibited stability up to 100 1C, and the
degradation of the samples usually started around
350 1C, with the whole mass degraded to char at
500 1C and completely burned off at 650 1C.

2.6. Castor oil polymers


Epoxidized castor oil has been used for the
preparation of interpenetrating polymer networks
(IPN), and these were characterized for their
dynamic mechanical behavior [57]. It was observed
that the cross-linked IPNs from the epoxidized oil
and adducts of tung oil with maleic anhydride had
very good compatibility. The hydroxyl groups of
epoxidized castor oil form hydrogen bonds with the
carbonyl groups in tung oil. These hydroxyls are
more reactive towards tung oil adducts than
epoxidized cottonseed oil. Yagci and coworkers

ARTICLE IN PRESS
V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

1004

70

Epoxy Conversion (%)

60
50
40
30
20
10

Neat ENLO-100
1% DEGDVE in ENLO-100
5% DEGDVE in ENLO-100
20% DEGDVE in ENLO-100

0
-10
0

50

100
150
200
Irradiation Time (sec.)

250

300

Fig. 23. The effect of the reactive diluents, diethylene glycol


divinyl ether (DEGDVE), on the epoxy ring-opening polymerization of sample ENLO-100 in the presence of 4 weight percent
(4-octyloxyphenyl) phenyl iodonium hexauoroantimonate
(OPPI) as a function of the irradiation time. Reprinted with the
permission from J Polym Sci: Part A: Polym Chem 2003; 41: 3451
r Wiley Periodicals, Inc. [50].

70

Epoxy Conversion (%)

60

ate fatty acid ester into esters of dimer and oligomer


acids. A xed bed reactor was used for the
oligomerization of dehydrated methylricinoleate.
The product was characterized by infrared, 1HNMR and mass spectroscopy. The kinetics of the
oligomerization of methyl ester of dehydrated
castor oil fatty acid was compared with the thermal
oligomerization process [59]. The catalytic reaction
followed second-order kinetics, whereas the thermal
process followed rst-order kinetics.
Schwank et al. [60] reported the copolymerization
of dehydrated castor oil with styrene. The polymerization was carried out in benzene and the
product was isolated for characterization. It was
observed that the copolymerization was very
difcult when the concentration of dehydrated
castor oil is more than 20%. Ashraf and coworkers
[61] prepared a blend of dehydrated castor oil and
epoxy resin. The miscibility of the two components
was examined by viscometric and ultrasonic techniques. Ashraf and coworkers [62] also studied the
miscibility of the blends of epoxidized dehydrated
castor oil and poly (methyl methacrylate). The
compatibility was investigated by differential scanning calorimetry and SEM.
2.7. Polymers from other oils

50
40
30
20

Neat ENLO-100
2 % DEGDEE in ENLO-100
6 % DEGDEE in ENLO-100
20 % DEGDEE in ENLO-100

10
0
-10
0

50

100
150
200
Irradiation Time (sec.)

250

300

Fig. 24. The effect of non-reactive diluents, di-(ethylene glycol)


diethyl ether (DEGDEE), on the epoxy ring-opening polymerization of sample ENLO-100 in the presence of 4 weight percent
(4-octyloxyphenyl) phenyl iodonium hexauoroantimonate
(OPPI) as a function of the irradiation time. Reprinted with the
permission from J Polym Sci: Part A: Polym Chem 2003; 41: 3451
r Wiley Periodicals, Inc. [50].

prepared styrenated castor oil and linseed oil by the


macromer technique [45].
Kumar and co-workers [58] oligomerized castor
oil over molybdenum oxide on a silicaalumina
support. They directly converted the methylricinole-

Drying and semidrying oils such as sesame,


sunower, safower, walnut oil have also been
investigated for the preparation of polymer by
different methods. Homopolymerization is not
favored due to the steric hindrance of the bulky
oil moieties. Triglyceride oil-based macromers of
sunower oil were prepared in two successive steps
[46]. First, a partial glyceride was prepared by
glycerolysis of sunower oil in the presence of
calcium oxide (CaO). Subsequently, a macromer
was prepared by transesterication of the partial
glyceride with methylmethacrylate in the presence of
the same catalyst. The prepared macromer was
copolymerized with styrene in the presence of
benzoyl peroxide, yielding a styrenated product. In
this case, a product with longer polystyrene
segments was obtained.
Kusefoglu and coworkers [63] studied the polymerization of high oleic sunower oil with different
comonomers. The effect of simultaneous addition of
bromine and acrylate to the fatty acid triglycerides
was studied. The yield for bromoacrylation of
sunower oil was 55%. The copolymers of sunower oil and styrene were formed via free radical

ARTICLE IN PRESS
V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

1005

Photolysis of Photoinitiator

[b]

[Ar 2 I + X + ] +

Ar 2 I + X +

[a]

ArI*

ArI

X*

ArI*

X*

+ R + H*X*

Initiation

[c]

H+ +

[d]

H+ +

H O+

H
O

O+

Propagation

[e]

H O+

H O
O+
O

[f]

[h]

O O
O H
O

Homopolymer (I)

[g]
O

+ H O

H
O+

O H

[i]

O+

[j]

O H
O
O

Homopolymer (II)
O

[L]

Scheme 7. The mechanism for the photoinitiated cationic polymerization of the epoxynorbornane linseed oil (ENLO). Reprinted with the
permission from J Polym Sci: Part A: Polym Chem 2003; 41: 3454 r Wiley Periodicals, Inc. [50].

mechanism. They also studied the polymerization of


acrylamide derivatives of fatty acid compounds [64].
The acrylamide functionality on the triglyceride of
sunower oil was introduced by the Ritter reaction.
3. Conclusion
In recent years, natural oils have become the
center of attraction for their potential use as starting
materials for the preparation of polymers. This is an

alternate route, which has the potential to augment


the use of petroleum-based polymers. These triglyceride oils mostly comprise unsaturated fatty acids,
and provide a wide scope for polymerization using a
variety of techniques. The unsaturation present in
these oils makes them ideal for the preparation of
bio-based polymers. Polymers prepared from
soybean oil have properties comparable with those
of conventional polymers. The mechanical properties of these polymers depend on the degree of

ARTICLE IN PRESS
V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

1006

Acknowledgements

100

Weight loss (%)

80

The authors are thankful to the chief editors of


this journal for their kind suggestions and editorial
corrections.

C87LIN30-ST28-DVB42 (S1)
C87LIN40-ST24-DVB36 (S2)
C87LIN50-ST20-DVB30 (S3)

60

C87LIN60-ST12-DVB28 (S4)
C87LIN70-ST08-DVB22 (S5)

40

References
20
0
0

100

200

300

400

500

600

700

Temperature (C)

Fig. 25. The temperature dependence of the weight loss during


the thermogravimetric analysis of different samples prepared
from 87% conjugated linseed oil (C87LIN), styrene (ST)
and divinylbenzene (DVB). Reprinted with the permission
from Biomacromolecules 2005; 6: 805 r American Chemical
Society [55].

cross-linking; increased cross-linking increases the


ultimate strength and decreases the elongation at
break. The dynamic mechanical properties of some
natural oil polymers suggest that they are ideal
replacements for petroleum-based polymers, i.e.,
some natural oil derived polymers possess very good
damping and shape-memory properties over a wide
range of temperature. The fatty acid ester groups
directly attached to the polymer backbone are
presumed to be responsible for damping properties.
Damping materials have numerous applications in
the aircraft, automobile, and machinery industries
for the reduction of unwanted noise as well as the
prevention of vibrational fatigue failure. Shapememory materials have applications in civil construction, mechanics and manufacturing, electronics
and communications, printing and packaging,
medical equipment, recreation and sports, and
household items.
The polymerization of these oils is carried out via
free radical and cationic polymerization reactions.
Different research groups in the world are studying
the properties of natural oils and their composites
for utilization as polymers, resins, varnishes and
paints. The fossil-based monomers are harmful to
environment. These are non-renewable as they are
derived mainly from petroleum-based materials.
The fossil-based feedstocks are depleting very
rapidly. Thus, the main goal for the researchers in
coming years is to produce viable polymers from the
natural resources.

[1] Khot SN, Lascala JJ, Can E, Morye SS, Williams GI,
Palmese GR, et al. Development and application of
triglyceride-based polymers and composites. J Appl Polym
Sci 2001;82(3):70323.
[2] Gunstone F. Fatty acid & lipid chemistry. New York:
Blackie Academic & Professional; 1996.
[3] Cunningham A, Yapp A. Liquid polyol compositions. US
Patent, 3,827,993, 1974.
[4] Bussell GW. Maleinized fatty acid esters of 9-oxatetracyclo4.4.1.2,5O1,6O8,10 undecan-4-ol. US Patent, 3,855,163,
1974.
[5] Hodakowski LE, Osborn CL, Harris EB. Polymerizable
epoxide-modied compositions. US Patent, 4,119,640, 1975.
[6] Trecker DJ, Borden GW, Smith OW. Method for curing
acrylated epoxidized soybean oil amine compositions. US
Patent, 3,979,270, 1976.
[7] Trecker DJ, Borden GW, Smith OW. Acrylated epoxidized
soybean oil amine compositions and method. US Patent,
3,931,075, 1976.
[8] Salunkhe DK, Chavan JK, Adsule RN, Kadam SS. World
oilseeds: chemistry, technology and utilization. New York:
Van Nostrand Reinhold; 1992.
[9] Force CG, Starr FS. Vegetable oil adducts as emollients in
skin and hair care products. US Patent, 4,740,367, 1988.
[10] Li F, Hanson MV, Larock RC. Soybean oildivinylbenzene
thermosetting polymers: synthesis, structure, properties and
their relationships. Polymer 2001;42:156779.
[11] Li F, Larock RC. New soybean oilstyrenedivinylbenzene thermosetting copolymers I: synthesis and characterization. J Appl Polym Sci 2001;80:65870.
[12] Li F, Larock RC. New soybean oilstyrenedivinylbenzene
thermosetting copolymers II: dynamic mechanical properties. J Polym Sci B Polym Phys 2000;38:272138.
[13] Li F, Larock RC. New soybean oilstyrenedivinylbenzene
thermosetting copolymers III: tensile stressstrain behavior.
J Polym Sci B Polym Phys 2001;39:6077.
[14] Li F, Larock RC. New soybean oilstyrenedivinylbenzene
thermosetting copolymers IV: good damping properties.
Polym Adv Technol 2002;13:43649.
[15] Li F, Larock RC. New soybean oilstyrenedivinylbenzene thermosetting copolymers V: shape-memory effect.
J Appl Polym Sci 2002;84:153343.
[16] Li F, Hou J, Zhu W, Zhang X, Xu M, Luo X, et al.
Crystallinity and morphology of segmented polyurethanes
with different soft-segment length. J Appl Polym Sci
1996;62:6318.
[17] Kim BK, Lee SY, Xu M. Polyurethanes having shapememory effects. Polymer 1996;37:5781.
[18] Li F, Zhang X, Hou J, Xu M, Luo X, Ma D, et al. Studies
on thermally stimulated shape-memory effect of segmented
polyurethanes. J Appl Polym Sci 1997;64:15116.

ARTICLE IN PRESS
V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008
[19] Kim BK, Lee SY, Lee JS, Baek SH, Choi YJ, Xu M.
Polyurethane ionomers having shape-memory effects. Polymer 1998;39:28038.
[20] Li F, Larock RC. New soybean oilstyrenedivinylbenzene
thermosetting copolymers VI: Timetemperaturetransformation cure diagram and the effect of curing conditions on
the thermoset properties. Polym Int 2003;52:12632.
[21] Gillham JK. Formation and properties of thermosetting and
high Tg polymeric materials. Polym Eng Sci 1986;26:142933.
[22] Enns JB, Gillham JK. Timetemperaturetransformation
(TTT) cure diagram: modeling the cure behavior of
thermosets. J Appl Polym Sci 1983;28:256791.
[23] Nunez L, Taboada J, Fraga F, Nunez MR. Kinetic study
and timetemperaturetransformation cure diagram for an
epoxy-diamine system. J Appl Polym Sci 1997;66:137788.
[24] Hazer B, Demirel SI, Borcakli M, Eroglu MS, Cakmak M,
Burak E. Free radical crosslinking of unsaturated bacterial
polyesters obtained from soybean oily acids. Polym Bull
2001;46:38994.
[25] Hamurcu EE, Baysal BM. Interpenetrating polymer networks of poly (dimethylsiloxane): 1. Preparation and
characterization. Polymer 1993;34:51637.
[26] Li F, Larock RC, Marks DW, Otaigbe JU. Fish oil
thermosetting polymers: synthesis, structure, properties and
their relationships. Polymer 2000;41:792539.
[27] Gruger GH. In: Stansby ME, editor. Fish oils: their
chemistry, technology, stability, nutritional properties and
uses. Connecticut: The AVI Publishing Company; 1967.
p. 330.
[28] Li F, Larock RC, Otaigbe JU. Fish oil thermosetting
polymers: creep and recovery behavior. Polymer 2000;41:
484962.
[29] Marks DW, Li F, Pacha CM, Larock RC. Synthesis of
thermoset plastics by lewis acid initiated copolymerization of
sh oil ethyl esters and alkenes. J Appl Polym Sci
2001;81:200112.
[30] Li F, Perrenoud A, Larock RC. Thermophysical and
mechanical properties of novel polymers prepared by the
cationic copolymerization of sh oils, styrene and divinylbenzene. Polymer 2001;42:1013345.
[31] Li F, Hasjim J, Larock RC. Synthesis, structure and
thermophysical and mechanical properties of new polymers
prepared by the cationic copolymerization of corn oil,
styrene and divinylbenzene. J Appl Polym Sci 2003;90:
18308.
[32] Gunstone FD. Industrial uses of soybean oil for tomorrow,
special report1996. Ames, IA: Iowa State University and
The Iowa Soybean Promotion Board; 1995.
[33] Kinabrew RG. In Tung oil in Mississippi: the competitive
position of the industry. MS: University of Mississippi; 1952.
[34] Li F, Larock RC. Thermosetting polymers from cationic
copolymerization of tung oil: synthesis and characterization.
J Appl Polym Sci 2000;78:104456.
[35] Li F, Larock RC. Synthesis, structure and properties of
new tung oilstyrenedivinylbenzene copolymers prepared
by thermal polymerization. Biomacromolecules 2003;4:
101825.
[36] Boelhouwer C, Klassen WA, Waterman HI. Res Corresp
Suppl Res (Lond) 1954;7:S62.
[37] Rheineck AE, Austin AO. In: Myers Raymond R, Long JS,
editors. Drying oilsmodication and use: treatise on
coatings part II, vol. 1. New York: M. Dekker; 1967.

1007

[38] Trumbo DL, Mote BE. Synthesis of tung oildiacrylate


copolymers via the DielsAlder reaction and properties of
lms from the copolymers. J Appl Polym Sci 2001;
80:236975.
[39] Conte LS, Lerekar G, Capella P, Catena M. Linseed oil
composition. Riv Ital Sost Gras 1979;56:33942.
[40] Thames SF, Wang Z, Brister EH, Hariharan R, King CL,
Panjanani KG. Internally plasticized and low VOC latex
compositions and applications thereof. US Patent, 6,
624,223, 2003.
[41] Tortorello AJ, Montgomery E, Chawla CP. Radiationcurable compositions comprising oligomers having an alkyd
backbone. US Patent, 6,638,616, 2003.
[42] Motawie AM, Hassan FA, Manich A, Aboul-Fetouh ME,
El-din A, Fakhr. Some epoxidized polyurethane and
polyester resins based on linseed oil. J Appl Polym Sci
1995;55:172532.
[43] Meneghetti SMP, de Souza RF, Monteiro AL, de Souza
MO. Subtitution of lead catalysts by zirconium in the
oxidative polymerization of linseed oil. Prog Org Coat
1998;33:21924.
[44] Tuman SJ, Chamberlain D, Scholsky KM, Soucek MD.
Differential scanning calorimetry study of linseed oil cured
with metal catalysts. Prog Org Coat 1996;28:2518.
[45] Gultekin M, Beker U, Guner FS, Erciyes AT, Yagci Y.
Styrenation of castor oil and linseed oil by macromer
method. Macromol Mater Eng 2000;283:1520.
[46] Akbas T, Beker UG, Guner FS, Erciyes AT, Yagci Y.
Drying and semidrying oil macromonomers III: styrenation
of sunower and linseed oils. J Appl Polym Sci 2003;
88:23736.
[47] Guner FS, Usta S, Erciyes AT, Yagci Y. Styrenation of
triglyeride oils by macromonomer technique. J Coat Technol
2000;72:10710.
[48] Hewitt DH, Armitage F. Manufacture of interpolymers of
styrene with polyhydric alcoholic mixed esters and of coating
compositions obtained therefrom. US Patent 2,586,652,
1952.
[49] Mallegol J, Lemaire J, Gardette JL. Drier inuence on the
curing of linseed oil. Prog Org Coat 2000;39:10713.
[50] Boquillon N, Frignant C. Polymer networks derived from
curing of epoxidised linseed oil: inuence of different
catalysts and anhydride hardeners. Polymer 2000;41:
860313.
[51] Zong Z, Soucek MD, Liu Y, Hu J. Cationic photopolymerization of epoxynorbornane linseed oils: the effects of
diluents. J Polym Sci A Polym Chem 2003;41:344056.
[52] Rajaraman SK, Powers WA, Crivello JV. Interaction of
epoxy and vinyl ethers during photoinitiated cationic
polymerization. J Polym Sci A Polym Chem 1999;37:
400718.
[53] Rajaraman SK, Powers WA, Crivello JV. Novel hybrid
monomers bearing cycloaliphatic epoxy and 1-propenyl
ether groups. Macromolecules 1999;32:3647.
[54] Zong Z, He J, Soucek MD. UV-curable organicinorganic
hybrid lms based on epoxynorbornane linseed oils. Prog
Org Coat 2005;53:8390.
[55] Turri B, Vicini S, Margutti S, Pedemonte E. Calorimetric
analysis of the polymerization process of linseed oil.
J Thermal Anal Calorim 2001;66:3438.
[56] Kundu PP, Larock RC. Novel conjugated linseed oilstyrenedivinylbenzene copolymers prepared by thermal

ARTICLE IN PRESS
1008

V. Sharma, P.P. Kundu / Prog. Polym. Sci. 31 (2006) 9831008

polymerization 1: effect of monomer concentration on the


structure and properties. Biomacromolecules 2005;6:
797806.
[57] Yin Y, Yao S, Zhou X. Synthesis and dynamic mechanical
behavior of crosslinked copolymers and IPNs from vegetable oils. J Appl Polym Sci 2003;88:18402.
[58] Kumar VG, Venkatachalam S, Rao KVC. Insitu dehydrooligomerization of castor oil, fatty acid ester into esters of
dimer and oligomer acids over molybdenum oxide on silicaalumina catalyst. J Polym Sci Polym Chem 1984;22:
231727.
[59] Kumar VG, Venkatachalam S, Rao KVC. Kinetics of
oligomerization of methyl ester of dehydrated castor oil fatty
acid over molybdenum oxide on silica-alumina catalyst in
comparision with the thermal oligomerization process.
J Polym Sci Polym Chem 1984;22:380514.

[60] Cassidy PE, Schwank. Copolymerization of dehydrated


castor oil with styrene: determination of reactivity ratios.
J Appl Polym Sci 1974;18:251726.
[61] Ashraf SM, Ahmad S, Riaz U, Alam M, Sharma HO.
Compatibility studies on dehydrated castor oil blend with
poly(methacrylic acid). J Macromol Sci A Pure Appl Chem
2005;42:140921.
[62] Ashraf SM, Ahmad S, Riaz U, Sharma HO. Studies on
miscibility of dehydrated castor oil epoxy blend with poly
(methyl methacrylate). J Appl Polym Sci 2006;100:3094100.
[63] Eren T, Kusefoglu SH. Synthesis and polymerization of the
bromoacrylated plant oil triglycerides to rigid, ame
retardant polymers. J Appl Polym Sci 2004;91:270010.
[64] Eren T, Kusefoglu SH. Synthesis and polymerization of the
acrylamide derivatives of fatty compounds. J Appl Polym
Sci 2005;97:226472.

You might also like